Mechanistic Studies of the Hydroamination of Norbornene with

Nov 14, 2008 - Jennifer L. McBee, Alexis T. Bell and T. Don Tilley*. Departments of Chemistry and Chemical Engineering, University of California, Berk...
0 downloads 7 Views 454KB Size
Published on Web 11/14/2008

Mechanistic Studies of the Hydroamination of Norbornene with Electrophilic Platinum Complexes: The Role of Proton Transfer Jennifer L. McBee,†,‡ Alexis T. Bell,‡,§ and T. Don Tilley*,†,‡ Departments of Chemistry and Chemical Engineering, UniVersity of California, Berkeley, Berkeley, California 94720, and Chemical Sciences DiVision, Lawrence Berkeley National Laboratory, 1 Cyclotron Road, Berkeley, California 94720 Received April 23, 2008; E-mail: [email protected]

Abstract: Hydroaminations of norbornene with arylsulfonamides and weakly basic anilines were achieved using electrophilic Pt(II) bis(triflate) complexes of the type L2Pt(OTf)2 (L2 ) tBu2bpy, tBuC6H4Nd C(CH3)C(CH3)dNC6H4tBu, (C6H5)2PCH2CH2P(C6H5)2, (C6F5)2PCH2CH2P(C6F5)2, S-BINAP). Pseudo-firstorder kinetics reveal little to no dependence of the reaction rate on the ancillary ligand. Mechanistic studies do not favor an olefin coordination mechanism but are instead consistent with a mechanism involving sulfonamide coordination and generation of an acidic proton that is transferred to the norbornene. It is postulated that the resulting norbornyl cation is then attacked by free sulfonamide, and loss of proton from this adduct completes the hydroamination. The platinum-sulfonamide complex readily undergoes deprotonation to give a µ-amido platinum-bridged dimer that was isolated from the reaction solution. These studies also involve use of Me3SiPh and Me3SnPh as non-nucleophilic proton traps. Cleavage of the Ph-E bonds was used to detect the acidic, catalytically active species.

Introduction

Scheme 1. Olefin Activation Mechanism for Hydroamination

A longstanding goal of research in catalysis is the development of general catalysts for the selective hydroamination of unactivated alkenes.1,2 The search for such catalysts has generated a number of plausible mechanistic strategies, and notable progress has been made for certain substrates. Specifically, hydroaminations have been observed for vinylarenes,3-5 dienes,6 alkynes,7,8 and electron-deficient alkenes.5,9-11 In addition, several reports of hydroaminations with weakly basic amines (such as sulfonamides, carboxamides, and electrondeficient anilines) have appeared.12-17 Despite these important advances, general catalysts that operate for unactivated alkenes †

Department of Chemistry, UC Berkeley. ‡ Lawrence Berkeley National Laboratory. § Department of Chemical Engineering, UC Berkeley. (1) Mu¨ller, T. E.; Beller, M. Chem. ReV. 1998, 98, 675–704. (2) Hartwig, J. F. Pure Appl. Chem. 2004, 507–516. (3) Utsunomiya, M.; Kuwano, R.; Kawatsura, M.; Hartwig, J. F. J. Am. Chem. Soc. 2003, 125, 5608–5609. (4) Utsunomiya, M.; Hartwig, J. F. J. Am. Chem. Soc. 2004, 126, 2702– 2703. (5) Kawatsura, M.; Hartwig, J. F. J. Am. Chem. Soc. 2000, 122, 9546– 9547. (6) Lo¨ber, O.; Kawatsura, M.; Hartwig, J. F. J. Am. Chem. Soc. 2001, 123, 4366–4367. (7) Doye, S. Synlett 2004, 10, 1653–1672. (8) Bytschkov, I.; Doye, S. Eur. J. Org. Chem. 2003, 935–946. (9) Fadini, L.; Togni, A. A. Chimia 2004, 58, 208–211. (10) Fadini, L.; Togni, A. Chem. Commun. 2003, 30–31. (11) Munro-Leighton, C.; Blue, E. D.; Gunnoe, T. B. J. Am. Chem. Soc. 2006, 128, 1446–1447. (12) Wang, X.; Widenhoefer, R. A. Organometallics 2004, 23, 1649–1651. (13) Taylor, J. G.; Whittall, N.; Hii, K. K. Org. Lett. 2006, 8, 3561–3564. (14) Liu, X.-Y.; Li, C.-H.; Che, C.-M. Org. Lett. 2006, 8, 2707–2720. (15) Karshtedt, D.; Bell, A. T.; Tilley, T. D. J. Am. Chem. Soc. 2005, 127, 12640–12646. 16562

9

J. AM. CHEM. SOC. 2008, 130, 16562–16571

and more basic amines, such as aniline and ammonia, have yet to be discovered. Only lanthanide catalysts exhibit a somewhat wide substrate scope covering a variety of alkene and amine substrates; however, these catalysts are highly moisture- and oxygen-sensitive.18,19 Much of the effort directed toward development of general hydroamination catalysts has centered on late transition metals, since catalysts based on these metals are expected to be more moisture- and oxygen-tolerant. Also, these metals offer a number of potential mechanisms for activation of the amine and alkene substrates. In particular, mechanisms based on nucleophilic attack of amine onto a coordinated alkene (Scheme 1) and amine activation by N-H oxidative addition (Scheme 2) have received (16) Bender, C. F.; Widenhoefer, R. A. J. Am. Chem. Soc. 2005, 127, 1070– 1071. (17) Zhang, J.; Yang, C.-J.; He, C. J. Am. Chem. Soc. 2006, 128, 1798– 1799. (18) Hong, S.; Marks, T. J. Acc. Chem. Res. 2004, 37, 673–686. (19) Roesky, P. W.; Mu¨ller, T. E. Angew. Chem., Int. Ed. 2003, 42, 2708– 2710. 10.1021/ja8030104 CCC: $40.75  2008 American Chemical Society

Hydroamination by Electrophilic Platinum Complexes Scheme 2. Amine Activation Mechanism for Hydroamination

considerable attention.7,20-24 Hydroaminations via the former mechanism would seem to require an electrophilic metal center in the catalyst, and for this reason research has focused on cationic complexes of the late transition metals. Several research groups have found that cationic late metal complexes with weakly coordinating anions catalyze a range of C-H or N-H additions across an alkene.22,25 In our laboratories, electrophilic platinum catalysts such as (COD)Pt(OTf)2, where COD ) 1,5cyclooctadiene, and [(η2-C2H4)PtCl2]2/AgBF4 have been found to mediate hydroarylations26 and hydroaminations15 of alkenes. The observed hydroaminations involved only weakly basic aniline and sulfonamide derivatives, and this reaction is associated with a substrate limitation in which amines with conjugate acid pKa values above ca. 4 are unreactive. There are several possible mechanistic explanations for this limitation; however, the pKa cutoff suggests the involvement of a proton transfer. The possible role of Brønsted acidic species in observed, metal-catalyzed hydroarylations and hydroaminations is further suggested by the fact that strong acids are known to be catalysts for such reactions.27-31 Thus, a detailed characterization of the role of electrophilic metal species in catalytic element-hydrogen additions to unsaturated substrates must include a thorough understanding of the potential participation of protons. In principle, this issue may be difficult to probe, since a standard mechanistic test for involvement of a Brønsted acid involves addition of a base (a proton trap). However, for reactions involving Lewis acidic metal centers and relatively acidic intermediates, an added base may inhibit the formation of products for a variety of reasons. Thus, it is important to develop mechanistic probes that reveal possible processes by which protonic species might result from the interaction of a Lewis acidic metal center with hydroamination substrates (an alkene (20) Senn, H. M.; Deubel, D. V.; Blo¨chl, P. E.; Togni, A. THEOCHEM 2000, 506, 233–242. (21) Senn, H. M.; Blo¨chl, P. E.; Togni, A. J. Am. Chem. Soc. 2000, 122, 4098–4107. (22) Hahn, C. Chem. Eur. J. 2004, 10, 5888–5889. (23) Brunet, J.-J.; Chu, N. C.; Diallo, O. Organometallics 2005, 24, 3104– 3110. (24) Brunet, J.-J.; Cadena, M.; Chu, N. C.; Diallo, O.; Jacob, K.; Mothes, E. Organometallics 2004, 23, 1264–1268. (25) Chianese, A. R.; Lee, S. J.; Gagne´, M. R. Angew. Chem., Int. Ed. 2007, 46, 4042–4059. (26) Karshtedt, D.; Bell, A. T.; Tilley, T. D. Organometallics 2004, 23, 4169–4171. (27) Schlummer, B.; Hartwig, J. F. Org. Lett. 2002, 4, 1471–1474. (28) Rosenfeld, D. C.; Shekhar, S.; Takemiya, A.; Utsunomiya, M.; Hartwig, J. F. Org. Lett. 2006, 8, 4179–4182. (29) Anderson, L. L.; Arnold, J.; Bergman, R. G. J. Am. Chem. Soc. 2005, 127, 14542–14543. (30) Dubovyk, I.; Watson, I. D. G.; Yudin, A. K. J. Am. Chem. Soc. 2007, 129, 14172–14173. (31) Cochran, B. M.; Michael, F. E. J. Am. Chem. Soc. 2008, 130, 2786– 2792.

ARTICLES Chart 1

and an amine). One possible reaction of this type was recently reported by Szuromi and Sharp, who described the interaction of (COD)Pt(OTf)2 with norbornene, which results in alkene coupling at the metal center and the elimination of triflic acid.32 Clearly, acidic species may also result from coordination of hydrogen-substituted amines to a metal center. This contribution describes a mechanistic investigation of hydroaminations catalyzed by electrophilic bis(triflate) platinum complexes. This research involves kinetic and mechanistic studies, including the use of Me3SiPh and Me3SnPh as trapping reagents for protons. The results point to the platinum-mediated generation of an acidic species, which plays a crucial role in the catalysis. Results and Discussion

The bis(triflate) platinum complex (COD)Pt(OTf)2 has previously been reported to catalyze hydroarylations26 and hydroaminations15 of alkenes. For these reactions, experimental evidence suggests that the COD ligand remains coordinated to platinum and does not directly participate in the catalysis. Thus, to probe the role of the platinum center in hydroaminations, it was of interest to vary the nature of the chelating, ancillary L2 ligand in catalysts of the type L2Pt(OTf)2. Synthesis and Characterization of L2Pt(OTf)2 Complexes.

These complexes were obtained in high yield by reaction of the corresponding L2PtCl2 compounds with 2 equiv of AgOTf in dichloromethane.33-35 In this manner, a range of triflate derivatives containing both phosphine- and nitrogen-based ligands were prepared (Chart 1). Because these complexes are moisture-sensitive and readily form µ-OH bridged dimeric complexes upon exposure to water, the synthetic procedures involved the rigorous exclusion of air.36,37 The L2Pt(OTf)2 complexes were completely characterized, and the nature of the L2-Pt interactions was investigated by single crystal X-ray crystallography. For each complex investigated, both triflate ligands are coordinated to platinum through one oxygen atom. The Pt-O distances (Å) vary depending on the ligand sets, from 2.060(4) and 2.063(4) for 1 (Figure 1) to 2.115(3) and 2.119(3) (32) Szuromi, E.; Sharp, P. R. Organometallics 2006, 25, 558–559. (33) Oberbeckmann-Winter, N.; Morse, X.; Braunstein, P.; Welter, R. Inorg. Chem. 2005, 44, 1391–1403. (34) Nilsson, P.; Puxty, G.; Wendt, O. F. Organometallics 2006, 25, 1285– 1292. (35) Liang, C.-C.; Lin, J.-M.; Lee, W.-Y. Chem. Commun. 2005, 19, 2462– 2464. (36) Strukul, G.; Varagnolo, A.; Pinna, F. J. Mol. Catal. A 1997, 117, 413– 423. (37) Bunkan, N. M.; Gagne´, M. R. Organometallics 2002, 21, 4711–4717. J. AM. CHEM. SOC.

9

VOL. 130, NO. 49, 2008

16563

ARTICLES

McBee et al.

Figure 1. ORTEP diagram of the X-ray crystal structure of 1. Hydrogen atoms are omitted for clarity. Bond lengths (Å): Pt-O4 ) 2.060(4), Pt-O3 ) 2.063(4), Pt-N1 ) 1.967(5), Pt-N2 ) 1.980(5).

with both electron-deficient and electron-rich functional groups being tolerated (Table 1, entries 7-15). A similar dependence on the conjugate acid pKa of the amine has been observed with other late transition metal catalysts.15 Variation of the catalyst employed did not reveal a change in the observed pKa limitation for the hydroamination. The exclusive formation of the exo products in these reactions indicates the intermediacy of a norbornyl cation, since insertions of norbornene into metal-ligand bonds result in endo products.42,43 To further evaluate involvement of a norbornyl cation in the hydroamination catalysis, norbornene and 4-tert-butylbenzenesulfonamide-d2 were subjected to catalytic conditions with 5 mol % 1 as the catalyst. By 2H NMR spectroscopy, four C-D resonances were observed for the hydroamination product. The large number of C-D resonances strongly implies participation of a norbornyl cation (only one C-D product would be expected for an insertion mechanism). Lastly, when 5 was employed as the catalyst, a 50:50 mixture of enantiomers was observed as the product (by HPLC Chiracel OD). The complete lack of enantioselectivity with 5 as the catalyst suggests that the reaction does not involve the intermediacy of diastereomers and is consistent with the involvement of a norbornyl cation. Finally, compounds 1-5 were observed to catalyze the isomerization of norbornene to tricyclo(2.2.1.0(2,6))heptane, which is known to occur via an intermediate norbornyl cation.42-44 For 1 as catalyst, an equilibrium of the two isomers was established after 1 day at 25 °C (Keq ) 1.1(7), eq 1).

complex (COD)Pt(OTf)2 is a catalyst for the hydroamination of norbornene with sulfonamides.15 For catalysts 1-5, hydroaminations occurred in near quantitative yields at 60 °C within 12 h. With catalyst 1, the amine scope for the hydroamination was explored (Table 1). With 5 mol % of 1 relative to norbornene, no hydroamination products were observed for morpholine, aniline, or p-anisidine (conjugate acid pKa > 4) at 100 °C over 3 days (Table 1, entries 1-3). However, if the conjugate acid pKa value for the amine substrate is lowered to below 4, as for pentafluoroaniline, p-nitroaniline, and 3,5-(bistrifluoromethyl)aniline, hydroamination proceeds at 50-80 °C in >95% yields, with high selectivity for the exo product (Table 1, entries 4-6). With more acidic amines (sulfonamides and benzamide with pKa < 0),39-41 the hydroaminations occurred

Ancillary Ligand Effects on Reaction Rate. The influence of the ancillary ligand set on hydroamination reaction rates was investigated with kinetic studies. Under pseudo-first-order conditions with excess norbornene (20 equiv) and sulfonamide (20 equiv), kobs values for each catalyst were determined at 37 °C. As shown in Table 2, there is only a small variation in rate constants among the catalysts employed, with nitrogen-based ancillary ligands (1, 2) giving somewhat higher rate constants than phosphine ligands (3-5). As a reference, the kobs value for triflic acid (5 mol %), a known catalyst for hydroaminations with sulfonamides, is two to 3 orders of magnitude greater than those for the platinum catalysts. Kinetic Studies on Hydroamination. Mechanistic studies of the hydroamination of norbornene with sulfonamide included attempts to establish a rate law using kinetic measurements. For this purpose, catalyst 1 was employed, and reactions were monitored at 37 °C in chloroform-d. First-order dependence of the reaction rate on the concentration of 1 was established under pseudo-first-order conditions of excess norbornene and sulfonamide (10-100 equiv). The reaction order in sulfonamide was determined by varying its concentration (0.01-0.1 M) in the presence of a controlled amount of 1 (0.005 M) and excess norbornene (0.2 M). Under these pseudo-first-order conditions, plots of [sulfonamide] versus time were linear over 2-3 halflives, and a plot of ln rate versus ln[sulfonamide] provided a reaction order of 1.1 for the sulfonamide. Variation of the

(38) Brunkan, N. M.; White, P. S.; Gagne´, M. R. Organometallics 2002, 21, 1565–1575. (39) Virtanen, P. O.; Maikkula, M. Tetrahedron Lett. 1968, 9, 4855–4858. (40) Laughlin, R. G. J. Am. Chem. Soc. 1967, 89, 4268–4271. (41) Gross, K. C.; Seybold, P. G. J. Org. Chem. 2001, 66, 6919–6925.

(42) Hughes, R. P.; Powell, J. J. Organomet. Chem. 1971, 30, C45–C47. (43) Gallazzi, M. C.; Hanlon, T. L.; Vitulli, G.; Porri, L. J. Organomet. Chem. 1971, 33, C45–C46. (44) Saunders, M.; Jarret, R. M.; Pramanik, P. J. Am. Chem. Soc. 1987, 109, 3735–3739.

Figure 2. ORTEP diagram of the X-ray crystal structure of 5 · C4H10O.

Hydrogen atoms and the C4H10O molecule are omitted for clarity. Bond lengths (Å): Pt-O2 ) 2.115(3), Pt-O1 ) 2.119(3), Pt-P5 ) 2.2258(14), Pt-P6 ) 2.2165(14).

for 5 (Figure 2) and 2.120(4) and 2.138(4) for 3,38 illustrating only small variations in the Pt-triflate interactions. Hydroamination with Weakly Basic Anilines and Sulfonamides. Previous studies found that the platinum bis(triflate)

16564

J. AM. CHEM. SOC.

9

VOL. 130, NO. 49, 2008

Hydroamination by Electrophilic Platinum Complexes

ARTICLES

Table 1. Intermolecular Hydroamination of Norbornenea

a Reactions were conducted with 0.1 mmol of substrate and norbornene and 5 mol % of 1 in 1 mL of ClCH2CH2Cl. b Yields were determined by 1H NMR spectroscopy using an internal standard and confirmed by GC-MS, and only the exo product was observed. c Norbornene and starting amine were isolated. d The long reaction time may be attributed to the low solubility of benzamide under the reaction conditions.

Table 2. Rate Constants for Catalytic Hydroamination of Norbornene

catalyst

kobs (s-1)

(tBu2bpy)Pt(OTf)2 (1) [tBuC6H4NdC(CH3)(CH3)CdNC6H4tBu]Pt(OTf)2 (2) [[(C6H5)2PCH2CH2P(C6H5)2]Pt(OTf)2 (3) [[(C6F5)2PCH2CH2P(C6F5)2]Pt(OTf)2 (4) [S-(BINAP)]Pt(OTf)2 (5) triflic acid

1.4(9) × 10-3 3.2(8) × 10-3 5.3(9) × 10-4 5.3(9) × 10-4 2.9(7) × 10-4 2.0(9) × 10-1

norbornene concentration (0.03-0.08 M) in the presence of 1 (0.005 M) and excess sulfonamide (0.2 M) gave only small changes in the reaction rate, and an analogous kinetic analysis gave a reaction order of 0.2 for norbornene, over the concentration range examined (see Supporting Information, Figure S7).

Under conditions employed for the kinetic studies, the rate of norbornene isomerization is much slower than the rate for hydroamination ( 2σ(I)] 334 0.037; 0.069; 0.072 0.788 3.026 -1.092

P212121 4 1.650 g/cm3 ω (0.3° per frame) total 24173 unique 8109 (Rint ) 0.0486) Lorentz polarization Abs (Tmax ) 0.913, Tmin ) 0.881) full-matrix least-squares 8109 [I > 2σ(I)] 613 0.031; 0.062; 0.063 1.018 1.070 -0.429

ensure no further changes in the concentrations. For 9: 1H NMR (CDCl3, 500 MHz): δ 9.10 (d, J ) 6.4 Hz, 2H, ArH), 8.17 (d, J ) 8.4 Hz, 2H, SO2ArH 9), 8.02 (s, 2H, ArH), 7.86 (overlapping with resonance for SO2ArH, 4H, ArH), 7.76 (d, J ) 8.4 Hz, 2H, SO2ArH), 4.92 (br s, 2H), 1.51 (s, 36H, CH3C overlapping with resonance for the CH3C of 1), 1.40 (s, 9H, CH3C). For 1: 1H NMR (CDCl3, 500 MHz): δ 8.53 (d, J ) 6.4 Hz, 2H, ArH), 7.98 (s, 2H, ArH), 7.79 (d, J ) 6.4 Hz, 2H, ArH), 1.51 (s, 36H, CH3C overlapping with resonance for the CH3C of 9). For 4-tertbutylbenzenesulfonamide: 1H NMR (CDCl3, 500 MHz): δ 7.86 (overlapping m, 4H, SO2ArH and ArH, 9), 7.58 (d, J ) 8.4 Hz, 2H, SO2ArH), 4.92 (br s, 2H), 1.35 (s, 9H, CH3CArSO2). 195Pt NMR (CD2Cl2, 85.9 MHz): δ -2010, -1530. 19F NMR (CD2Cl2, 376.5 MHz): δ -75.9, -78.0. [(tBu2bpy)Pt(COD)](OTf)2 (10). To 0.050 g of 1 (0.066 mmol) was added a solution of 0.010 g of COD (0.092 mmol) in 5 mL of CH2Cl2. The reaction mixture was stirred for 2 h. The solvent was then removed under reduced pressure and the resulting solid was washed three times with 5 mL of pentane. The yellow powder was dried under reduced pressure for 24 h (95% yield). 1H NMR (CDCl3, 500 MHz): δ 8.25 (d, J ) 4 Hz, 2H), 8.10 (s, 2H), 7.77 (d, J ) 4 Hz, 2H), 6.33 (br t, 4H), 3.12 (d, J ) 6.8 Hz, 4H), 2.62 (d, J ) 6.8 Hz, 4H), 1.47 (s, 18H). 19F NMR (CDCl3, 376.5 MHz): δ -79.2. 13C{1H} NMR (CDCl3, 125.8 MHz): δ 167.2, 157.0, 150.9, 128.3, 124.9, 119.9, 36.3, 30.1. Anal. Calcd for C28H36N2F6O6PtS2: C, 38.66; H, 4.17; N, 3.22 Found: C, 38.51; H, 4.11; N, 2.94. 4-tert-butylbenzenesulfonamide-d2. To 1.00 g of 4-tert-butylbenzenesulfonamide in 5 mL of diethyl ether was added 3 mL of D2O. The reaction mixture was left to stir for 24 h. The solvent was removed under reduced pressure, and the resulting product was treated three more times with the ether/D2O mixture. After the last removal of solvent, the product was dissolved in diethyl ether and 1.00 g of MgSO4. After 4 h, the solution was filtered, and the solvent was removed under reduced pressure (96% yield, 95% deuterium incorporation). 1H NMR (CDCl3, 500 MHz): δ 7.87 (d, J ) 6.8 16570

J. AM. CHEM. SOC.

9

5

VOL. 130, NO. 49, 2008

6

C59H78Cl2F6N6O10Pt2S4 1734.6 141 yellow, block 0.05 × 0.04 × 0.03 mm3 triclinic 2.58 < 2θ < 49.54 a ) 11.779(3) Å b ) 16.296(4) Å c ) 20.316(8) Å R ) 100.69(3)° β ) 100.84(3)° γ ) 96.14(2)° V ) 3613.7(19) Å3 P1j 2 1.594 g/cm3 ω (0.3° per frame) total 20696 unique 12158 (Rint ) 0.0158) Lorentz polarization Abs (Tmax ) 0.886, Tmin ) 0.820) full-matrix least-squares 12158 [I > 2σ(I)] 811 0.039; 0.109; 0.114 1.035 2.414 -1.005

Hz, 2H), 7.55 (d, J ) 6.8 Hz, 2H), 1.35 (s, 9H). 13C{1H} NMR (CD2Cl2, 125.8 MHz): δ 156.6, 139.0, 126.1, 126.0, 30.7. 2H NMR (CDCl3, 76.7 MHz): δ 4.61 (s). General Procedure for Catalytic Runs. Reactions occurred in 5 mm NMR tubes equipped with J. Young Teflon screw caps. Temperature-controlled oil baths were utilized for heating. A solution of 0.106 mmol of olefin and 0.106 mmol of amine was prepared in 1 mL of 1,2-dichloroethane. To this solution was added 0.005 mmol of catalyst. The progress of the reaction was monitored by 1H NMR spectroscopy (0.05 mL cyclohexane-d12 was added in order to obtain a lock signal). Once no further spectral change was observed by NMR, the solution was passed through silica to isolate the substrate or product. Identities of the organic products were confirmed by 1H NMR spectroscopy59,60 and GC-MS. All kinetic data was collected on a Bruker DRX-500 MHz spectrometer and an internal standard, 5.0 mg (0.0250 mmol) of bis(p-fluorophenyl)methane, was used. For kinetic runs, the sample was placed in an NMR probe preheated to 310.4 K and calibrated using an external methanol standard.61 Single-scan spectra were obtained using an automated acquisition program that was started immediately after placing the sample in the probe, and the peaks were integrated relative to the internal standard. General Considerations for X-ray Structure Determinations. The X-ray analyses of compounds 1, 5, and 6 were carried out at UC Berkeley CHEXRAY crystallographic facility. Measurements were made on Bruker SMART CCD area detector with graphitemonochromated Mo KR radiation (λ ) 0.71073 Å). Data were integrated by the program SAINT and analyzed for agreement using XPREP. Empirical absorption corrections were made using SAD(59) Ackermann, L.; Kaspar, L. T.; Gscheri, C. J. Org. Lett. 2004, 6, 2515– 2518. (60) Fleming, I.; Frackenpohl, J.; Ila, H. J. Chem. Soc., Perkin Trans. 1 1998, 1229–1236. (61) Ammann, C.; Meier, P.; Merbach, A. E. J. Magn. Reson. 1982, 46, 319–321.

Hydroamination by Electrophilic Platinum Complexes

ABS. Structures were solved by direct methods and expanded using Fourier techniques. All calculations were performed using the SHELXTL crystallographic package. Selected crystal structure and refinement data can be found in Table 3. X-ray Structure Determination of 1. X-ray quality crystals of 1 formed upon slow vapor diffusion of ether into a methylene chloride solution of 1 at -35 °C. The compound crystallizes in the monoclinic space group I2/a with one molecule in the asymmetric unit. X-ray Structure Determination of 5. X-ray quality crystals of 5 formed upon slow vapor diffusion of ether into a methylene chloride solution of 5 at -35 °C. The compound crystallizes in the orthorhombic chiral space group P212121 with one molecule of ether and one molecule of 5 in the asymmetric unit. The enantiomorph of the space group and enantiomer of the molecule were determined by inspection of Friedel pairs of reflections followed by comparison of least-squares refinements. The results were unequivocally in favor of the enantiomorph reported. X-ray Structure Determination of 6. X-ray quality crystals of 6 were isolated from a catalytic run by slow vapor diffusion of

ARTICLES

ether into the catalytic run solution of methylene chloride at -35 °C. The compound crystallizes in the triclinic space group P1j with two molecules in the asymmetric unit and one molecule of methylene chloride. Acknowledgment. We thank Herman van Halbeek for NMR assistance and Dr. Fred Hollander and Dr. Allen Oliver for assistance with X-ray crystallography. This work was supported by the Director, Office of Energy Research, Office of Basic Energy Sciences, Chemical Sciences Division of U.S. Department of Energy under Contract No. DE-AC02-05CH11231. Supporting Information Available: Kinetic analysis, text describing complete X-ray experimental details and X-ray crystallographic data for 1, 5, and 6 in CIF format. This material is available free of charge via the Internet at http://pubs.acs. org. JA8030104

J. AM. CHEM. SOC.

9

VOL. 130, NO. 49, 2008

16571